Você está na página 1de 11

The Cauchy-Davenport Theorem for Finite Groups

Paul Balister
Department of Mathematical Sciences, University of Memphis
Memphis, TN, USA
pbalistr@memphis.edu

Jeffrey Paul Wheeler


Department of Mathematics, University of Pittsburgh
Pittsburgh, PA, USA
jeffreypaulwheeler@hotmail.com

Abstract

The Cauchy-Davenport theorem states that for any two nonempty subsets A and
B of Z/pZ we have |A + B| ≥ min{p, |A| + |B| − 1}, where A + B := {a + b
mod p | a ∈ A, b ∈ B}. We generalize this result from Z/pZ to arbitrary finite
(including non-abelian) groups. This result from early in 2006 is independent of
Gyula Károlyi’s1 2005 result in [13] and uses different methods.

1. Motivation

The problems we will be considering lie in the area of Additive Number


Theory. This relatively young area of Mathematics is part of Combinatorial
Number Theory and can best be described as the study of sums of sets of
integers. As such, we begin by stating the following definition:
Definition 1.1. [Sumset]
For subsets A and B of a group G, define
A + B := {a + b | a ∈ A, b ∈ B}
where + is the group operation2.

We note that originally G = Z/pZ but much work (including this one) has
been done and is being done in arbitrary groups.
1The authors wish to thank Gyula for introducing them to this problem and encouraging
work on it. Regrettably we did not let Gyula know that we were making progress, hence the
independent results. We discovered Gyula had a result the day before this work was presented to
the Combinatorics seminar at Memphis.
2We are not suggesting that G is abelian, but rather being consistent with the notation for
the sumset in cases where G is Z/pZ, Z, or an abelian group. Later we will introduce the more
appropriate notation.
1
2

A simple example of a problem in Additive Number Theory is given two


subsets A and B of a set of integers, what facts can we determine about
A + B? We will state a result regarding this example shortly. Note that a
very familiar problem in Number Theory, namely Lagrange’s theorem that
every nonnegative integer can be written as the sum of four squares, can be
expressed in terms of sumsets. In particular,
Theorem 1.2. [Lagrange’s Four Square Theorem]
Let N0 = {x ∈ Z | x ≥ 0} and let S = {x2 | x ∈ Z}. Then
N0 = S + S + S + S.

As well the the binary version of Goldbach’s Conjecture can be restated in


terms of sumsets.
Theorem 1.3. [Goldbach’s Conjecture]
Let E = {2x | x ∈ Z, x ≥ 2} and let P = {p ∈ Z | p is prime }. Then
(1) E ⊆ P + P.

In other words, every even integer is greater than 2 is conjecture to be the


sum of two primes. Notice that we do not have set equality in equation (1)
because 2 ∈ P.

2. Background

The theorem we wish to extend was first proved by Augustin Cauchy in


18133 [3] and later independently reproved by Harold Davenport in 1935 [5]
(Davenport discovered in 1947 [6] that Cauchy had previously proved the
theorem). In particular,
Theorem 2.1. [Cauchy-Davenport]
Let k, l be positive integers. If A, B ⊆ Z/pZ, p prime, with |A| = k ≤ p and
|B| = l ≤ p, then |A + B| ≥ min{p, k + l − 1} where A + B := {a + b | a ∈
A and b ∈ B}.

We note that in 1935 Inder Chowla [4] extended the result to composite
moduli m when 0 ∈ B and the other members of B are relatively prime
to m. As well it is worth noting that in 1996 Alon, Nathanson, and Ruzsa
provided a simple proof of this theorem using the Polynomial Method[1].

Of interest to this work is Gyula Károlyi’s extension of the theorem to


abelian groups[9],[10]. Before we state the theorem, though, a useful defini-
tion:
3Cauchy used this theorem to prove that Ax2 + By 2 + C ≡ 0(mod p) has solutions provided
that ABC 6≡ 0. This is interesting in that Lagrange used this result to establish his four squares
theorem.
3

Definition 2.2 (Minimal Torsion Element).


Let G be a group. We define p(G) to be the smallest positive integer p for
which there exists a nonzero element g of G with pg = 0 (or, if multiplicative
notation is used, g p = 1). If no such p exists, we write p(G) = ∞.
Lemma 2.3.
The p in Definition 2.2 is the smallest prime factor of |G| provided G is
finite.

Proof.
Suppose |G| = p1 e1 · p2 e2 · · · · · pn en where p1 < p2 < · · · < pn are primes
end the ei are positive integers. By Cauchy’s Theorem there is an element
g ∈ G such that g p1 = 1. Suppose there were a smaller prime q such that
there were a gc ∈ G where gcq = 1. Then | h gc i | = q and by Lagrange’s
Theorem q||G|. This is a contradiction.

Equipped with Definition 2.2 we state


Theorem 2.4. (Károlyi[9],[10])
If A and B are nonempty subsets of an abelian group G, then |A + B| ≥
min{p(G), |A| + |B| − 1} where A + B := {a + b|a ∈ A, b ∈ B}.

Again, our goal is to extend this result to arbitrary finite groups. A necessary
tool will be the famous and very useful result:
Theorem 2.5 (Feit-Thompson[8]).
Every group of odd order is solvable.

3. A Basic Structure of Finite Solvable Groups

Throughout this section G will be a finite solvable group, i.e. there exists a
chain of subgroups
{1} = G0 E G1 E G2 E · · · E Gn = G
such that Gi /Gi−1 is abelian for i = 1, 2, 3, . . . , n.

By definition, there is some K = Gn−1 E G such that G/K = H where H


is abelian. Pick a representative ehi ∈ G for each coset hi = K e hi ∈ H. So
for each g ∈ G, there is a ki ∈ K and there is an hi ∈ H (in particular,
the coset representative) such that g = kie hi . Given this, we build a useful
structure for finite solvable groups. First, define
(2) ψH : G → K × G/K = K × H by ψ(g) = (ki , h̃i ).
4

(Note well that the second coordinate is the coset representative.) As well,
define an operation ? on K × H by
(3) (k1 , h˜1 ) ? (k2 , h˜2 ) := (k1 φh˜1 (k2 )ηh˜1 ,h˜2 , h˜1 h˜2 ).
where
(4) φh̃ : K → Aut(K)
in particular, φh̃ (k) = h̃k h̃−1 , and
(5) ηh˜1 ,h˜2 = h˜1 · h˜2 · (h
]1 h2 )
−1
∈K
with h̃ the coset representative of h in G/K 4. Notice that η : H × H → K
(think cosets instead of coset representatives). Later examples will illustrate
that this η plays an analogous role to “carrying the 1” in addition of real
numbers.
Lemma 3.1 (A Basic Structure of Solvable Groups).
Let G be a solvable group with K E G. Upon fixing the coset representatives
in H = G/K, ψH in (2) is an isomorphism from G to the group (K × H, ?).

Proof.
Since we have fixed the coset representatives h = h̃ for H, for every g ∈ G
there exists a unique k ∈ K such that g = kh; i.e. ψH is one-to-one and
onto. Suppose g1 = k1 h1 and g2 = k2 h2 . Then
(6) ψH (g1 ) ? ψH (g2 ) = (k1 , h1 ) ? (k2 , h2 )
(7) = (k1 φh1 (k2 )ηh1 ,h2 , h1 h2 )
−1
(8) = (k1 h˜1 k2 h˜1 h˜1 h˜2 (h
] −1
1 h2 ) , h1 h2 )
−1
(9) = ψH (k1 h˜1 k2 (h˜1 h˜1 )h˜2 (h
] −1 ]
1 h2 ) h1 h2 )

(10) = ψH (k1 h˜1 k2 h˜2 )


(11) = ψH (k1 h1 k2 h2 )
= ψH (g1 g2 )
¤

In summary, for A ⊆ G, we have an isomorphism A → K × H, in particular,


A∼ = {(k1 , h1 ), (k2 , h2 ), . . . , (kt , ht )} for some k1 , k2 , . . . , kt ∈ K and (fixed)
h1 , h2 , . . . ht ∈ H. We note that it is certainly not the case that the ki ’s nor
the hj ’s are distinct.

It is worth noting that the construction of ? on K × H is more general than


the semi-direct product. Indeed, G may not be a semi-direct product of K
and H.
4i.e. for each h ∈ H = G/K there exists h̃ ∈ G such that h = K e
h.
5

Before we continue, two illustrative examples.


Example 3.2.
Let Q be the quaternion group, namely Q = {±1, ±i, ±j, ±k} with Q’s mul-
tiplication table stated for easy reference in Table 1 (Note: multiplication is
row · column). Put K = {±1, ±k} and since |Q/K| = 2,
{1} E K E Q;
i.e. Q is a solvable group.

Table 1. Multiplication Table for the Quaternion Group Q

· 1 i j k
1 1 i j k
i i −1 k −j
j j −k −1 i
k k j −i −1

So Q/K = {K, Kj} and we choose 1 as our coset representative of K and


j as the coset representative of Kj (see Table 2).

Table 2. Cosets of Q and Their Representatives

Cosets of Q/K Representative


K = {±1, ±k} 1
Kj = {±j, ±i} j

Hence the order of Kj in Q/K is 2 however the order of j in Q is 4. This


means
(12) ηj,j := j̃ · j̃ · (jg
· j)−1
= j̃ · j̃ · 1̃−1 (note that the coset representative of −1 is 1)
= j · j · 1 (now we multiply as in G)
= −1.

As well
(13) ηj,1 = η1,j
:= 1̃ · j̃ · (1g
· j)−1
= 1̃ · j̃ · j̃ −1
= 1 · j · j −1
= 1 · j · −j
= 1.
6

And clearly

(14) η1,1 = 1.

Before continuing with the example, we list the elements of Q as written


using the structure of Lemma 3.1 with K = {±1, ±k} in Table 3. Note that
the first component is in K and the second is either of the selected coset
representatives 1 or j.

Table 3. Elements of Q Written as in the Basic Structure with


Coset Representatives as in Table 2

q∈Q 1 −1 i −i j −j k −k
(k, h) (1, 1) (−1, 1) (−k, j) (k, j) (1, j) (−1, j) (k, 1) (−k, 1)

Thus, since i = −k · j,

(15) i·i∼
= ψH (i) ? ψH (i)
(16) = (−k, j) ? (−k, j) (see table 3)
−1
(17) = (−k{j(−k)j · ηj,j }, jj)
(18) = (−k{−i(−j)(−1)}, j 2 )
(19) = (−k · −k, 1)
(the multiplication in the second slot is as coset multiplication)
(20) = (−1, 1)

= −1.

Which is what we hoped for since i · i = −1.

To show we were not just lucky,

(21) i·k ∼
= ψH (i) ? ψH (k)
(22) = (−k, j) ? (k, 1) (see table 3)
−1
(23) = (−k{j(k)j · ηj,1 }, j1)
(24) = (−kjk(−j)1, j)
(25) = ([kj]2 , j)
(26) = (−1, j)

= −j.

and
7

(27) j·i∼
= ψH (j) · ψH (i)
(28) = (1, j)(−k, −j) (see table 3)
−1
(29) = (1j(−k)j · ηj,−j }, j(−j)
(30) = (j(−k)(−j)(−1), j(j))
(31) = (−jkj, 1)
(the multiplication in the second slot is as coset multiplication)
(32) = (−k, 1)
∼ −k.
=

¤
Example 3.3.
Let p be a prime. Then
¡ ¢
Z/p2 Z ∼
= pZ/p2 Z × Z/pZ, ?
where H = {0, 1, . . . , p − 1} ∼
= Z/pZ which we will write as {0, 1, . . . , p − 1}
and K = {0, p, . . . , (p − 1)p} ∼
= pZ/p2 Z which we will write as {0, p, . . . , (p−
1)p}.

Hence
Z/p2 Z ∼
= {(0, 0), (0, 1), . . . , (0, p−1), (p, 0), . . . , (p, p−1), . . . , ([p−1]p, p−1)}
.

Table 4. Elements of Z/p2 Z Written as in the Basic Structure

Z/p2 Z 0 1 ··· p − 1 p p + 1 ··· p2 − 1


2
(pZ/p Z, Z/pZ) (0, 0) (0, 1) · · · (0, p − 1) (p, 0) (p, 1) · · · ([p − 1]p, p − 1)

Hence
(33)
3 + [p2 − 2] ∼
= (0, 3) + ([p − 1]p, [p − 2])
(34) = (0 + φ3 ([p − 1]p) + η3,[p−2] , 3 + [p − 2])
(35) = ({3 + [p − 1]p + [p2 − 3]} + {3 + [p − 2] + [3 + p − 2]−1 }, p + 1)
(36) = (−p + 3 + [p − 2] + 1−1 , 1)
(37) = (1 + 1−1 , 1)
(38) = (0, 1)

=1
8

Before leaving this section, we note that (as stated earlier) neither Z/p2 Z
nor the quaternion group is the semidirect product of its respective K and
H.

Before proceeding, developing some notation will be helpful.


Definition 3.4.
For G a finite solvable group, we have K = Gn−1 E G. Putting H = G/K
and for S ⊆ G,
(39) S=∼ {(ki , hi ) where ki ∈ K and hi ∈ H}.
We will define
S 1 := {ki ∈ K|∃hi ∈ H such that (ki , hi ) ∈ S} and
S 2 := {hi ∈ G\K|∃ki ∈ K where (ki , hi ) ∈ S}.

In other words, S 1 is the collection of first coordinates of S and S 2 is the


collection of second coordinates of S when S is written as in (39).

4. The Cauchy-Davenport Theorem for Finite Solvable Groups

Let G be a solvable group and let S and T be subsets of G. Put s = |S| and
t = |T |. As previously stated, there exists a K E G so that H = G/K with
|H| = σ. Thus
S∼= {(ku , hi )} for some i ∈ {1, . . . , σ} where ku ∈ K for u ∈ {1, . . . , s}
T ∼
= {(kv , hj )} for some j ∈ {1, . . . , σ} where kv ∈ K for v ∈ {1, . . . , t}.
Hence
Definition 4.1.
Define S1 = {(kj1 , h1 )}, S2 = {(kj2 , h2 )}, . . . , Sα = {(kjα , hα )} where |S1 | =
s1 ≥ |S2 | = s2 ≥ · · · ≥ |Sα | = sα (thus 1 ≤ j1 ≤ s1 , 1 ≤ j2 ≤ s2 , etc.).
Construct T1 , T2 , . . . , Tβ in a similar manner.

Following Definition 4.1,


Remark 4.2.
We have S = S1 ∪ S2 ∪ · · · ∪ Sα and T = T1 ∪ T2 ∪ · · · ∪ Tβ , hence |S| = s =
s1 + s2 + · · · + sα and |T | = t = t1 + t2 + · · · + tβ .

Since we will be concerned with non-abelian groups,


Definition 4.3.
For an arbitrary group G with S, T ⊆ G,
S · T := {st|s ∈ S and t ∈ T }.
9

Since the second coordinates will be distinct, the set {(S1 · Tj )2 |1 ≤ j ≤ β}


will have β elements. But S 2 , T 2 ⊆ H, hence by Theorem 2.4 |S 2 · T 2 | ≥
α + β − 1. Thus
Remark 4.4.
Since α + β − 1 = (β) + (α − 1), there are at least α − 1 elements in the set
{(Si · Tj )2 |1 < i ≤ α, 1 ≤ j ≤ β}.
Lemma 4.5.
For each i ∈ {1, . . . , α} and each j ∈ {1, . . . , β},
|Si · Tj | = |(Si · Tj )1 | = |Si 1 φhi (Tj1 )ηhi ,hj | = |(Si )1 · (Tj )1 |

Proof.
Noting that the second coordinate is the same establishes the first equality.
The second equality is just the definition of the product. The final equality
holds since conjugation is an isomorphism as is multiplying by ηhi ,hj , which
is some fixed element in K (hi and hj are fixed).

¤
Theorem 4.6.
Suppose S, T ⊆ G, G solvable of order n with |S| = s, |T | = t and s + t − 1 <
p(G). Then |S · T | ≥ s + t − 1.

Proof.
We will proceed by induction on n, namely we will assume the theorem holds
for solvable groups of order less then n. We have that there exists a K E G
such that H = G/K. We will express S and T as in Definition 4.1 and we
choose S and T such that β ≥ α. Together with Remark 4.4 we get (since
there are at least α − 1 non-empty sets (Si · Tj ), 1 < i ≤ α, 1 ≤ j ≤ β)5
(40)
|S · T | ≥ |S1 · T1 | + |S1 · T2 | + · · · + |S1 · Tβ | + α − 1
By Lemma 4.5, we have
(41) = |S11 · T11 | + |S11 · T21 | + · · · + |S11 · Tβ1 | + α − 1
By the induction hypothesis on K which is solvable and of order < n, we get
(42) ≥ s1 + t1 − 1 + s1 + t2 − 1 + · · · + s1 + tβ − 1 + α − 1
(43) ≥ βs1 + t1 + t2 + · · · + tβ − β + α − 1
(44) = αs1 + t + (β − α)s1 − (β − α) − 1 (since β ≥ α)
(45) ≥ s + t + 0 − 1 (since s1 ≥ 1)
(46) = s + t − 1.

5By Remark 4.4, there are α − 1 second coordinates that come from these sets.
10

5. The Cauchy-Davenport Theorem for Finite Groups

We now extend Theorem 4.6 to all finite groups.


Theorem 5.1.
Let G be a finite group and let S, T ⊆ G with |S| = s and |T | = t. Then
|S · T | ≥ min{p(G), s + t − 1}.

Proof.
The case |S| = |T | = 1 is trivial. If G is of even order, then p(G) = 2. If G
is of odd order, then by Theorem 2.5, G is solvable. The result then follows
from Theorem 4.6.

6. A Related Problem

Very related to the Cauchy-Davenport Theorem is a conjecture of Paul Erdős


and Hans Heilbronn. In the early 1960’s they conjectured that if the sumset
addition in the theorem is restricted to distinct elements then the lower
bound changes slightly. In particular,
Theorem 6.1. [Erdős-Heilbronn Conjecture]
Let p be a prime and A, B ⊆ Z/pZ with A 6= ∅ and B =6 ∅. Then |A+̇B| ≥
min{p, |A| + |B| − 3}, where A+̇B := {a + b (mod p) | a ∈ A, b ∈ B and
a 6= b }.

The conjecture was first proved for the case A = B by J.A. Dias da Silva
and Y.O. Hamidounne in 1994 [7] using methods from linear algebra with
the more general case established by Noga Alon, Melvin B. Nathanson, and
Imre Z. Ruzsa using the polynomial method in 1995 [1]. Gyula Károlyi
extended this result to abelian groups for the case A = B in 2004 [10] and
to cyclic groups of prime powered order in 2005 [13].

What is interesting to note is how much more difficult the restricted addition
makes the problem. The Cauchy-Davenport Theorem was proven immedi-
ately but the Erdős-Heilbronn Conjecture was open for more than 30 years.
The authors of this paper have as well extended the conjecture of Erdős
and Heilbronn to Finite Groups [2] using similar techniques as in this paper.
The increased difficulty of the problem is represented well by requiring a
much stronger structure on finite solvable groups than what was used here.
Curious readers are encouraged to read J. Wheeler’s Ph.D. thesis [14].
11

References
[1] Alon, Noga and Nathanson, Melvyn B. and Ruzsa, Imre The polynomial method and re-
stricted sums of congruence classes, Journal of Number Theory, Volume 56, 1996, pgs. 404-417.

[2] Balister, Paul N. and Wheeler, Jeffrey Paul, The Erdős-Heilbronn problem for finite groups,
to appear in Acta Arithmetica.

[3] Cauchy, A. Recherches sur les nombres, J. École Polytech, Volume 9, 1813, pgs. 99-116.

[4] Chowla, Inder, A theorem on the addition of residue classes: application to the number Γ(k)
in Waring’s problem., Proceedings of the Indian Academy of Sciences, Section A, 1, (1935)
242–243.

[5] Davenport, H. On the addition of residue classes, Journal of the London Mathematical
Society, Volume 10, 1935, pgs. 30-32.

[6] Davenport, H., A historical note, Journal of the London Mathematical Society, 22, (1947)
100–101.

[7] Dias da Silva, J. A. and Hamidoune, Y. O., Cyclic spaces for Grassmann derivatives and
additive theory, The Bulletin of the London Mathematical Society, 26 No.2, (1994) 140–146.
[8] Feit, Walter and Thompson, John G. Solvability of groups of odd order, Pacific Journal of
Mathematics, Volume 13, 1963, pgs. 775-1029, Reviewer: M. Suzuki.

[9] Károlyi, Gyula On restricted set addition in abelian groups, Annales Universitatis Scientiarum
Budapestinensis de Rolando Eötvös Nominatae. Sectio Mathematica, 46 (2003) 47–53.

[10] Károlyi, Gyula The Erdős-Heilbronn problem in abelian groups, Israel Journal of Mathemat-
ics, 139 (2004) 349–359.

[11] Károlyi, Gyula The Cauchy-Davenport theorem in group extensions, L’ Enseignement


Mathematique, 51 (2005) 239–254.

[12] Károlyi, Gyula An inverse theorem for the restricted set addition in abelain groups, Journal
of Algebra, 290 (2005) 557–593.

[13] Károlyi, Gyula A compactness argument in the additive theory and the polynomial method,
Discrete Mathematics, 302 (2005) 124–144.

[14] Wheeler, Jeffrey Paul The Cauchy-Davenport theorem and the Erdős-Heilbronn problem for
finite groups, Ph.D. Thesis, http://jeffreypaulwheeler.com/, 2008.

Você também pode gostar